Category Theory
Daniel Naylor

Contents

1Definitions and Examples
2The Yoneda Lemma
3Adjunctions
4Limits
5Monads
6Filtered Colimits
7Regular Categories
Index

1 Definitions and Examples

Definition 1.1 (Category). A category C consists of:

  • (a)
    a collection ob C of objects A,B,C,.
  • (b)
    a collection mor C of morphisms f,g,h,.
  • (c)
    two operations dom , cod from mor C to ob C: we write f : A B for “f is a morphism and dom f = A and cod f = B”.
  • (d)
    an operation from ob C to mor C sending A to 1A : A A.
  • (e)
    a partial binary operation (f,g)fg on mor C, such that fg is defined if and only if dom f = cod g, and in this case we have dom fg = dom g and cod fg = cod f.

These are subject to the axioms:

  • (f)
    f1A = f and 1Ag = g when the composites are defined.
  • (g)
    f(gh) = (fg)h whenever fg and gh are defined.

Remark 1.2.

Example 1.3.

Definition 1.4 (Functor). Let C and D be categories. A functor F : C D consists of mappings F : ob C ob D and F + mor C mor D such that:

  • F(dom f) = dom Ff

  • F(cod f) = cod Ff

  • F(1A) = 1FA

  • F(fg) = (Ff)(Fg) whenever fg is defined.

We write Cat for the category of small categories and the functors between them.

Example 1.5.

Definition 1.6 (Natural transformation). Given categories C and D, and two functors CGFD, a natural transformation α : F G assigns to each A ob C a morphism αA : FA GA in D, such that for any AfB in C, the square

  FA       F B

FααGfABfGA       GB
commutes (we call this square the naturality square for α at f). Given α as above, and β : G H, we define βα : F H by (βα)A = βAαA. We write [C,D] for the category of functors C D and natural transformations between them.

Example 1.7.

We have isomorphisms of categories: e.g. F : Rel Relop defined by FA = A, FR = Ro = {(b,a)|(a,b) R} is its own inverse.

But we have a weaker notion of equivalence of categories.

Lemma 1.8. Assuming that:

Then α is an isomorphism in [C,D] if and only if αA is an isomorphism in D for each A.

Proof.

Definition 1.9 (Equivalence of categories). Let C and D be categories. An equivalence between C and D consists of functors F : C D and G : D C together with natural isomorphisms α : 1C GF, β : FG 1D. We write C D if there exists an equivalence between C and D.

We say P is a categorical property if

(C has P and C D)D has P.

Example 1.10.

Definition 1.11 (Faithful / full / essentially surjective). Let F : C D be a functor.

  • (a)
    We say F is faithful if, given f and g in mor C, (Ff = Fg, dom f = dom g, cod f = cod g)f = g.
  • (b)
    We say F is full if, for every g : FA FB in D, there exists f : A B in C with Ff = g.
  • (c)
    We say F is essentially surjective if, for any B ob D, there exists A ob C with FAB.

Note that if F is full and faithful, it’s essentially injective: given FA gFB in D, the unique AfB with Ff = g is an isomorphism.

We say D C is a full subcategory if the inclusion D C is a full functor.

Lemma 1.12. Assuming that:

  • F : C D

Then F is part of an equivalence CD if and only if F is full, faithful, essentially surjective.

Proof.

Definition 1.13 (Skeleton). By a skeleton of a category C, we mean a full subcategory containing just one object from each isomorphism class.

We say C is skeletal if it’s a skeleton of itself.

Example. Matk is a skeletal category; it’s isomorphic to the skeleton of fdVectk consisting of the spaces kn.

However, working with skeletal categories involves heavy use of the axiom of choice.

Definition 1.14 (Monomorphism / epimorphism). Let f : A B be a morphism in a category C. We say f is a monomorphism (or monic) if, given ChgA, fg = fhg = h. We say f is an epimorphism (or epic) if it’s a monomorphism in Cop.

We write AfB to indicate that f is monic, and AfB to indicate that it’s epic.

We say C is balanced if every arrow which is monic and epic is an isomorphism.

We will call a monic morphism e split if it has a left inverse (and similarly we may define the notion of split epic).

Example 1.15.

2 The Yoneda Lemma

Definition 2.1 (Locally small). We say a category C is locally small if, for any two objects A and B, the morphisms A B in C are parameterised by a set C(A,B).

If A is an object of a locally small category C, we have a functor C(A,) : C Set sending B to C(A,B) and a morphism BgC to the mapping (fgf) : C(A,B) C(A,C) (this is functorial since composition in C is associative).

Dually, we have C(,B) : Cop Set.

Lemma 2.2 (Yoneda). Assuming that:

Then

Proof.

Corollary 2.3. Assuming that:

Then AC(A,) is a full and faithful functor Cop [C,Set].

Proof. Substitute C(B,) for F in Lemma 2.2(i): we have a bijection from C(B,A) to the collection of natural transformations C(A,) C(B,).

For a given f, the natural transformation C(f,) sends g : B C to gf, so this is functorial by associativity of composition C.

Similarly, we have a full and faithful functor C [Cop,Set] sending A to C(,A). We call this the Yoneda embedding: it allows us to regard any locally small category C as a full subcategory of a Set-valued functor category.

Compare with Cayley’s Theorem in group theory (every group is isomorphic to a subgroup of a permutation group) and ‘Dedekind’s Theorem’ (every poset is isomorphic to a sub-poset of a power set).

Definition 2.4 (Representable). We say a functor F : C Set is representable if it’s isomorphic to a C(A,) for some A. By a representation of F, we mean a pair (A,x) where x FA is such that Φ(x) is an isomorphism. We call x a universal element of F.

Corollary 2.5. Suppose (A,x) and (B,y) are both representations of F. Then there is a unique isomorphism AfB such that (Ff)(x) = y.

Proof. (Ff)(x) = g is equivalent to saying that

    𝒞(B,∙)              𝒞(A, ∙)


𝒞ΦΦ(((f,yx∙)))         F
commutes, so f must be the unique isomorphism, whose image under Yoneda is Φ(x)1Φ(y).

Proof of Yoneda(ii).

Suppose for the moment that C is small, so that [C,Set] is locally small. Given two functors C ×[C,Set] Set: the first sends an object (A,F) to FA, and a morphism (AfA,FαF) to the diagonal of
  F A       F A′


FααFfAA′fF′ ′A       F ′A ′
The second is the composite

C ×[C,Set]Y × 1[C,Set]op ×[C,Set][C,Set](,)Set

where Y is a Yoneda embedding. Then Φ and Ψ define a natural isomorphism between these two.

In elementary terms, this says that if x FA, and x FA is its image under the diagonal, then Ψ(x) is the composite

C (A,)C(f,)C(A,)Ψ(x)FαF.

This makes sense without the assumption that C is small, and it’s true since the composite maps

1Af(Ff)(x)αA(Ff)(x).

Example 2.6.

In Set, products are just cartesian products (also in Gp, Rng, Top, …). coproducts in Set are disjoint unions A B = (A ×{0}) (B ×{1}). In Gp, coproducts are free products G H.

In Set, the equaliser of AgfB is the inclusion of {a A|f(a) = g(a)} and the coequaliser of (f,g) is the quotient of B by the smallest equivalence relation containing {(f(a),g(a))|a + A}.

Note that in Set, all monomorphisms and all epimorphisms are regular, but in Top, a monomorphism XfY is regular if and only if X is topologised as a subspace of Y . An epimorphism XfY is regular if and only if Y is topologised as a quotient of X.

Note that if f is both regular monic and regular epic, then it’s an isomorphism since the pair (g,h) of which its equaliser must satisfy g = h.

Warning. The following terminology is not standard. These are usually (both!) referred to as “generating”, but to avoid confusion, in this course we will refer to them with separate names.

Definition 2.7 (Separating / generating family). Let G be a family of objects of a locally small category C.

  • (a)
    We say G is a separating family if the functors C(G,), G G are jointly faithful, i.e. given a parallel pair AgfB, the equations fh = gh for all h : G A with G G imply f = g.
  • (b)
    We say G is a detecting family if the G(G,) jointly reflect isomorphisms, i.e. given AfB, if every GgB with G G factors uniquely through f, then f is an isomorphism.

If G = {G}, we call G a separator or a detector.

Lemma 2.8.

Proof.

Example 2.9.

By definition, the functors C(A,) : C Set preserve monomorphisms, but they don’t always preserve epimorphisms.

Definition 2.10 (Projective). We say an object P in a locally small category Cis projective if C(P,) preserves epimorphisms, i.e. if given

         P


fg Q      R
there exists h : P Q with gh = f. Dually, P is injective if it’s projective in Cop.

If P satisfies this condition for all g in some class E of epimorphisms, we call it E-projective.

In [C,Set], we consider the class of pointwise epimorphisms, i.e. those α such that αA is surjective for all A.

Corollary 2.11. functors of the form C(A,) are pointwise projective in [C,Set].

Proof. Immediate from Yoneda; given

           𝒞 (A, ∙)


αβ Q        R
with β pointwise epic, Φ(α) RA is βA(y) for some y QA, so βΨ(y) = α.

[C,Set] has enough pointwise projectives”:

Proposition 2.12. Assuming that:

Then there exists a pointwise epimorphism PF where P is pointwise projective.

Proof. Set P = (A,x)C(A,) where the disjoint union is over all pairs (A,x) with A ob C and x FA. A morphism P Q is uniquely determined by a family of morphisms C(A,) Q. . Hence P is pointwise projective, since all the C(A,) are. But we have α : P F whose (A,x)-th component is Ψ(x) : C(A,) F and this is pointwise epic since any x FA appears as Ψ(x)(1A).

3 Adjunctions

Definition 3.1 (Adjunction, D. Kan 1958). Let C and D be categories. An adjunction between C and D consists of functors F : C D and G : D C together with, for each A ob C and B ob D, a bijection between morphisms FA B in D and morphisms A GB in C, which is natural in A and B. (If C and D are locally small, this means that D(F,) and C(,G) are naturally isomorphic functors Cop ×D Set.)

We say F is left adjoint to G, or G is right adjoint to F, and we write (F G).

Example 3.2.

Theorem 3.3. Assuming that:

  • G : D C is a functor

  • for A ob C, let (A G) be the category whose objects are pairs (B,f) where B ob D and f : A GB, and whose morphisms (B,f) (B,f) are morphisms g : B B making
     A        GB


ffG′g        GB ′
    commute.

Then specifying a left adjoint for G is equivalent to specifying an initial object of (AG) for each A.

Proof. First suppose (FG). For each A ob C, let ηA : A GFA be the morphism corresponding to 1FA : FA FA. Then (FA,ηA) is an initial object of (AG): given any f : A GB, the diagram

  A        GF A

ηfGAg         GB
commutes if and only if g corresponds to f under the adjunction, by naturality of the adjunction bijection.

So there’s a unique morphism (FA,ηA) (B,f) in (AG).

Conversely, suppose given in initial object (FA,ηA) in (AG) for each A. We make F into a function C D: given AfB, Ff is the unique morphism (FA,ηA) (FB,ηBf) in (AG). Functoriality comes from uniqueness: given BgC, (Fg)(Ff) and F(gf) are both morphisms (FA,ηA) (FC,ηCgf) in (AG). The adjunction bijection sends AfGB to the unique morphism (FA,ηA) (B,f) in (AG), with inverse sending FAgB to (Gg)ηA : A GB. This is natural in A since η is a natural transformation 1C GF and natural in B since G is functorial.

Corollary 3.4. Assuming that:

Then there is a canonical natural isomorphism α : F F.

Proof. (FA,ηA) and (FA,ηA) are both initial in (AG), so there’s a unique isomorphism αA between them. α is natural: given AfB, (Ff)αA and αB(Ff) are both morphisms (FA,ηA) (FB,ηBf) in (AG), so they’re equal.

As a result of this, we will often talk about “the” left adjoint of a functor (when it exists), because we don’t usually care about which one in the isomorphism class we use.

Lemma 3.5. Assuming that:

  • CGFDKHE

  • (FG) and (HK)

Then (HFGK).

Proof. Given A ob C, C ob E, we have bijections between morphisms HFA C, morphisms FA KC, and morphisms A GKC which are both natural in A and C, D.

Corollary 3.6. Assuming that:

Then the square of left adjoints commutes up to natural isomorphism.

Proof. By Lemma 3.5, both ways round are left adjoint to HF = KG, so by Corollary 3.4 they’re isomorphic.

We saw in Theorem 3.3 that an adjoint (FG) gives rise to a natural transformation η : 1C GF, called the unit of the adjunction. Dually, we have 𝜀 : FG 1D, the counit of (FG).

Theorem 3.7. Assuming that:

Then specifying an adjunction (FG) is equivalent to specifying a natural transformation η : 1C GF and 𝜀 : FG 1D satisfying the two commutative diagrams:
 F         FGF            G        GF G


F1𝜀FηF        aFnd           η1GGG𝜀         G

Proof. Suppose (FG). We defined η in the proof of Theorem 3.3, and 𝜀 is defined dually. Since 𝜀FA corresponds to 1GFA, the composite 𝜀FA(FηA) corresponds to 1GFAηA = ηA. But by definition 1FA corresponds to ηA. The other identity is dual.

Conversely, suppose given η and 𝜀 satisfying the triangular identities. Given FAfB, we define Φ(f) = (Gf)ηA : A GFA GB. Dually, given AgGB, we define Ψ(g) = 𝜀B(Fg). Then ΨΦ(f) = Ψ((Gf)ηA) = 𝜀B(FGf)FηA = f(𝜀FA)(FηA) = f, and dually ΦΨ(g) = g. Naturality of Φ and Ψ follows from naturality of η and 𝜀.

In Definition 1.9, we had natural isomorphisms α : 1C GF and β : FG 1D. These look like the unit and counit of an adjunction (FG): do they satisfy the triangular identities? No, but we can always change them:

Proposition 3.8. Assuming that:

  • F : C D, G : D C, α : 1C GF and β : FG 1D be an equivalence of categories as defined in Definition 1.9

Then there exist isomorphisms α : 1C GF and β : FG 1D satisfying the triangular identities. In particular, (FGF).

Proof. We define α = α and take β to be the composite

FG(FGβ)1FGFG(Fα G)1FGβ1 D.

Note that FGβ = βFG, since

    FGF G      FG

FβββGβ FG         1
 FG            𝒟
commutes by naturality of β, and β is monic. Similarly, GFα = αGF.

To verify the triangular identities, consider

 F         FGF         FGF GF


           F           FGF

F1(F((1βαFβF−αFβFFFG1α)FG−)F1−)−11=(FGFα)−1       F
which commutes by naturality of β1.

For the second triangular identity, we have

  G        GF G        GF GF G


           G           GF G

α1(α((1GGGG−GGGβF1FβGα)−βG1))−−11= (αGFG)−1       G β
 G
Hence by Theorem 3.7 we have (FG). But (β)1 and α1 also satisfy the triangular identities for and adjunction (GF).

Lemma 3.9. Assuming that:

Then

Proof.

Definition 3.10 (Reflection). By a reflection, we mean an adjunction such that the right adjoint is full and faithful (equivalently: the counit is an isomorphism). We say D C is a reflective subcategory if it’s full and the inclusion D C has a left adjoint.

Example 3.11.

4 Limits

Definition 4.1 (Diagram). Let J be a category (almost always small, and often finite). By a diagram of shape J in a category C, we mean a functor D : J C. The objects D(j), j ob J are called vertices of D, and morphisms D(α), α mor J are called edges of D.

For example, if J is the category

∙      ∙

∙      ∙
a diagram of shape J is a commutative square in C.

If J is instead

∙      ∙


∙      ∙
then a diagram of shape J is a not-necessarily-commutative square.

Definition 4.2 (Cone, limit). Let D : J C be a diagram. A cone over D consists of an object A (its apex) together with morphisms λj : A D(j) for each j ob J (the legs of the cone) such that

            A

                        ′
λλDjj(′αD)(j)              D (j )
commutes for each α : j j in J.

A morphism of cones (A,(λj|j ob J)) (B,(μj|j ob J)) is a morphism f : A B such that μjf = λj for all j. We have a category Cone(D) of cones over D; a limit for D is a terminal object of Cone(D).

Dually, a colimit for D is an initial cone under D.

PIC

If Δ : C [J,C] is the functor sending A to the constant diagram with all vertices A then a cone over D is a natural transformation ΔA D.

Also, Cone(D) is another name for (ΔD), defined as in Theorem 3.3op.

So by Theorem 3.3, C has limits for all diagrams of shape J if and only if Δ has a right adjoint.

Example 4.3.

Proposition 4.4. Assuming that:

Then

Proof.

Definition 4.5 (Limit preserving / reflecting / creating). Let F : C D be a functor.

  • (a)
    We say F preserves limits of shape J if, given D : J C and a limit cone (L,(λj|j ob J)) for it, (FL,(Fλj|j ob J)) is a limit for FD : J D.
  • (b)
    We say F reflects limits of shape J if given D : J C, any cone over D which maps to a limit cone in D is a limit in C.
  • (c)
    We say F creates limits of shape J if, given D : J C and a limit cone (L,(λj|j ob J)) over FD, there exists a cone over D whose image under F is (L,(λj)), and any such cone is a limit in C.

We say a category C is complete if it has all small limits.

Corollary 4.6. In each of the statements of Proposition 4.4, we may replace ‘C has’ by either ‘D has and G : D C preserves’ or ‘C has and D C creates’.

Proof. Exercise.

Example 4.7.

Remark 4.8. In any category, AfB is monic if and only if

 A       A

11ffAAA       B
is a pullback. Hence, if D has pullbacks, then any monomorphism in [C,D] is pointwise monic, since its pullback along itself is contsructed pointwise.

Lemma 4.9. Assuming that:

Then G preserves all limits which exist in D.

Proof 1. Suppose (FG), and suppose C and D have limits of shape J. Then the diagram

    𝒞           𝒟


FΔΔ[J,F][J,”𝒞”]      [J,”𝒟”]
commutes, and all the functors in it have right adjoints, so
    [J,”𝒟 ”]     [J,”𝒞”]


[JliG,mGJ]𝒟           𝒞
commutes up to isomorphism by Corollary 3.6.

Proof 2. Suppose given D : J D and a limit cone (L,(λj|j ob J)) over it. Give a cone (A,(μj : A GD(j))) over GD, the transposes μj¯ : FA D(j) form a cone over D by naturality of the adjunction, so induce a unique μ¯ : FA L such that λjμ¯ = μj¯ for all j.

Then μ : A GL is the unique morphism satisfying (Gλj)μ = μj for all j.

Lemma 4.10. Assuming that:

Then for each A ob C, (AG) has limits of shape J and the forgetful functor (AG)UD creates them.

Proof. Suppose given D : J (AG); write D(j) = (UD(j),fj : A GUD(j)) and let (L,(λj|j ob J)) be a limit for UD. Since the edges of D are morphisms in (AG), the fj form a cone over GUD, so there’s a unique f : A GL satisfying (Gλj)f = fj for all j.

So (L,f) is the unique lifting of L to an object of (AG) which makes the λj into morphisms (L,f) (UD(j),fj) in (AG). The fact that these morphisms form a limit cone is straightforward.

Can we represent an initial object as a limit?

Lemma 4.11. Assuming that:

Then specifying an initial object of C is equivalent to specifying a limit for 1C : C C.

Proof. First suppose I is initial. The unique morphisms I A, A ob C, form a cone over 1C, and it’s a limit cone since if (A,(fB : A B|B ob C)) is any cone over 1C, then fI is its unique factorisation through the one with apex I.

Conversely, suppose given a limit (I,(fA : I A|A ob C)) for 1C. Then I is weakly initial (i.e. it admits morphisms to every object of C); and if g : I A then gfI = fA. In particular, fAfI = fA for all A, so fI is a factorisation of the limit cone through itself, so fI = 1I and I is initial.

The ‘primitive’ Adjoint Functor Theorem follows from Lemma 4.10, Lemma 4.11 and Theorem 3.3. But it only applies to preorders (see Example Sheet).

Theorem 4.12 (General Adjoint Functor Theorem). Assuming that:

Then G : D C has a left adjoint if and only if G preserves small limits and satisfies the solution-set condition: for every A ob C, there’s a set {(Bi,fi)|i I} of objects of (AG) which is collectively weakly initial.

Proof.

Example 4.13.

Definition 4.14 (Subobject). By a subobject of A ob C, we mean a monomorphism AA. We order subobjects by (AA) (AA) if there exists

A′              A′′


        A
We write Sub C(A) for this preorder.

We say C is well-powered if every Sub C(A) is equivalent to a small poset.

For example, Set is well-powered since the inclusions A A form a representative set of subobjects of A. It is well-copowered since isomorphism classes of epimorphisms AB correspond to equivalence relations on A.

Lemma 4.15. Assuming that:

Then k is monic.

Proof. Suppose given DmlP with kl = km. Then fhl = gkl = gkm = fhm, but f is monic so hl = hm. So l and m are both factorisations of

  D      A


hkllB
through the pullback, and hence l = m.

Theorem 4.16 (Special Adjoint Functor Theorem). Assuming that:

Then G : D C has a left adjoint if and only if it preserves all small limits.

Proof.

Example 4.17. Consider the inclusion KHausITop. Tychonoff’s Theorem says KHaus is closed under (small) products in Top. It’s closed under equalisers, since equalisers of pairs in KHaus are closed inclusions.

So KHaus is complete, and I preserves limits. KHaus and Top are locally small, and KHaus is well-powered since subobjects of X is isomorphic to inclusions of closed subspaces. And KHaus has a coseparator [0,1], by Uryson’s Lemma.

So by Theorem 4.16, I has a left adjoint β.

Remark 4.18.

5 Monads

Suppose we have CGFD, (FG). How much of the adjunction can we describe in terms of C (supposing we can’t know anything about D, or know very little about it)?

We have:

From the triangular identities of Theorem 3.7, we obtain the commutative triangles:

             T       TT          T       T T


(1) :      T1μηT        T(2) :      ηT1Tμ        T
and from naturality of 𝜀 we obtain
               T TT      TT


(3) :       TμμμμT T T       T

Definition 5.1 (Monad). A monad on a category C is a triple (T,η,μ) = 𝕋 where T : C C, and η : 1C T and μ : TT T satisfy the commutative diagrams

  T       TT


T1μηT        T
  T       TT

η1μTT        T
   TT T      TT


TμμμμT TT        T

Example 5.2.

Does every monad come from an adjunction?

Answered by Eilenberg-Moore and by Kleisli (1965).

Note that the monad of Example 5.2(a) is induced by SetUM × ()[M,Set] and that of Example 5.2(b) is induced by SetUPCSLatt, where CSLatt is the category of complete semilattices (posets, with arbitrary joins). The free complete semilattice on A is P(A): every f : A US extends uniquely to f¯ : P(A) S where f¯(A) = {f(a)|a A}.

An M-set (respectively a complete semilattice) is a set A equipped with a suitable mapping M × A A (respectively P(A)A).

Definition 5.3 (Eilenberg-Moore algebra). Let 𝕋 = (T,η,μ) be a monad on C. By an Eilenberg-Moore algebra for 𝕋 we mean a pair (A,α) where A ob C and α : TA A satisfies

            A        TA             TTA      T A

(4) :      ηA1Aα         A(5) :        TμαααA TA       A
A homomorphism f : (A,α) (B,β) is a morphism f : A B satisfying
              TA       TB


(6) :      Tαβff A        B
We write C𝕋 for the category of 𝕋-algebras and homomorphisms.

Proposition 5.4. Assuming that:

Then the forgetful functor C𝕋G𝕋C has a left adjoint F𝕋, and the adjunction induces the monad 𝕋.

Proof. We define F𝕋A = (TA,μA) (an algebra by (2) and (3)) and F𝕋(AfB) = Tf (a homomorphism by naturality of μ). Clearly, F𝕋 is functorial and G𝕋F𝕋 = T.

We establish the adjunction using Theorem 3.7: its unit is η, and the counit 𝜀(A,α) is just α (a homomorphism F𝕋A (A,α), by (5), and natural by (6)).

The triangular identity

  F 𝕋        F 𝕋G𝕋F 𝕋


1            F 𝕋
is just (1), and
  G 𝕋        G 𝕋F𝕋G 𝕋


1            G 𝕋
is (4).

Finally, G𝜀F𝕋A = μ by definition of F𝕋A. So the adjunction induces (T,η,μ).

Note: CGFD induces 𝕋, we can replace D by its full subcategory on objects FA.

So in trying to construct D, we may assume F is surjective (indeed, bijective) on objects. The morphisms FA FB in D must correspond to morphisms A GFB = TB in C.

Definition 5.5 (Kleisli category). Let 𝕋 be a monad on C. The Kleisli category C𝕋 is defined by ob C𝕋 = ob C, morphsims AfB in C𝕋 are morphisms AfTB in C. The identity AA is AηATA, and the composite of AfBgC is AfTBTgTTCμCTC.

For the unit and associative laws, consider the diagrams

  A       TB        TT B


fT1μηTBBB                 TB
  A         T A


  T B       T TB


μfTη1μAfTTBBB         T B
  A       TB        TTC        TT TD      T TD


fTTμTμμTμgTCμTDhDhDD                 TC         TT D       T D

Proposition 5.6. Assuming that:

Then there is an adjunction CG𝕋F𝕋C𝕋 inducing the monad 𝕋.

Proof. We define F𝕋A = A and F𝕋(AfB) = AfBηBTB. F𝕋 preserves identities by definintion, and preserves composition by

  A      B        TB


         C        TC       T TC

fg1gTηT1μfBgC1TCCC                        T C
We define G𝕋A = TA, and G𝕋(AfB) = TATfTTBμBTB. G𝕋 preserves identities by (1), and preserves composites by
  T A       TT B       T TTC      T TC


TTμTμμTμfTBμTCgCgCC         TB         T TC       T C
We verify the adjunction using Theorem 3.7: G𝕋F𝕋(f) = Tf by (1) so G𝕋F𝕋 = T and we take η as unit of the adjunction.

We define TA𝜀AA to be TA1TATA. To verify the naturality square

  TA       TB


F𝜀𝜀f𝕋ABAG𝕋f      B
the lower composite is TATfTTBμBTB and the upper one is TATfTTBμBTBηTBTTBμBTB, which agree since (2) tells us that μBηTB = 1B.

The triangular identities become

F𝜀𝕋FηAFA 𝕋A        F= GF A     FA        ηTηT1TηTATTATAAA        TTA        TTT A      TT A
and
ηGG𝜀GAAA         G=F GA     GA  η1ηTTTTAAA A       T TA      TA
Finally, G𝕋𝜀F𝕋A = μA, so (F𝕋G𝕋) induces the monad 𝕋.

Given a monad 𝕋 on C, we write Adj (𝕋) for the category whose objects are adjunctions (CGFD) inducing 𝕋, and morphisms (CGFD) (CGFD) are functors DKD satisfying KF = F and GK = G.

Theorem 5.7. The Kleisli adjunction (CC𝕋) is an initial object of Adj (𝕋), and the Eilenberg-Moore adjunction (CC𝕋 is terminal.

Proof. Suppose given (CGFD) in Adj (𝕋). We define K : D C𝕋 by KB = (GB,G𝜀B) (an algebra by one of the triangular identities for η and 𝜀, and naturality of 𝜀), K(BgB) = Gg (a homomorphism by naturality of 𝜀). K is functorial since G is, G𝕋K = G is obvious, and KFA = (GFA,G𝜀FA) = (TA,μA) = F𝕋A.

So K is a morphism of Adj (𝕋).

Suppose K : D C𝕋 is another such: then we must have KB = (GB,βB) where β : GFG G is a natural transformation since Kg = Gg is a homomorphism KB KB for all g : B B. Also, since KF = F𝕋, we have βFA = μA = G𝜀FA for all A.

For any B, we have naturality squares

     GF GF GB      GF GB


GGβGβGF𝜀F𝜀B𝜀GFGBB𝜀GBBBGF GB         GB
whose left edges are equal, and whose top edge is split epic, so we obtain G𝜀B = βB for all B. So K = K.

We define H : C𝕋 D by HA = FA and H(AfB) = FAFfFGFB𝜀FBFB. H preserves identities and satisfies HF𝕋 = F, by the first triangular identity for η and 𝜀.

H preserves the composite AfBgC by

  F A        FGF B         FGF GF C     F GF C


FF𝜀F𝜀𝜀F𝜀fGFGFFgFFB𝜀GCCFFgCC        FB            FGF C        F C
Also GHA = GFA = TA = G𝕋A and

GH(AfB) = (TATfTTBμBTB) = G𝕋(AfB).

So H is a morphism of Adj (𝕋). Note that H is full and faithful, since it sends AfGFB to its traspose across (FG).

If H : C𝕋 D is any morphism of Adj (𝕋), we must have HA = FA = HA for all A, and since GH = G𝕋 and the adjunctions have the same unit, H must send the transpose AfB of AfGFB to its transpose across (FG). So H = H.

C𝕋 has coproducts if C does, but has few other limits or colimits. In contrast, we have:

Proposition 5.8. Assuming that:

Then

Proof.

Given (CGFD), (FG), how can we tell when K : D C𝕋 is part of an equivalence?

Note: H : C𝕋 D is an equivalence if and only if F is essentially surjective.

We call (FF) (or the functor G) monadic if K : D C𝕋 is part of an equivalence.

Lemma 5.9. Assuming that:

  • CGFD is an adjunction inducing the monad 𝕋 on C

  • for every 𝕋 algebra (A,α), the pair FGFA𝜀FAFαFA has a coequaliser in D

Then K : D C𝕋 has a left adjoint L.

Proof. Write FAλ(A,α)L(A,α) for the coequaliser. For any homomorphism f : (A,α) (B,β) the two left hand squares in

    FGF A     F A        L (A, α)


F𝜀FFλFLF𝜀FλαAG(AffβB(BF,,fα)β)FGF B     F B        L (B, β)
commute, so we get a unique Lf making the right hand square commute. As usual, uniqueness implies functoriality of L.

For any B ob D, morphisms L(A,α) B correspond to morphisms FAfB satisfying f(Fα) = f𝜀FA. If f¯ : A GB is the transpose of f across (FG), then f(Fα) transposes to f¯α : GFA GB, whereas f𝜀FA transposes to Gf. But we can write f = 𝜀B(Ff¯) by the proof of Theorem 3.7, so Gf = (G𝜀B)(GFf¯). So f(Fα) = f𝜀FA if and only if

   GF A       GF GB


GαGfF𝜀fBA          GB
commutes, which happens if and only if f¯ : (A,α) KB in C𝕋.

Naturality of the bijection follows from that of ff¯.

Note that since G𝕋K = G, we have LF𝕋F by Corollary 3.6, and L preserves coequalisers.

Definition 5.10 (Reflexive / split coequaliser diagram / G-split).

  • (a)
    We say a parallel pair AgfB is reflexive if there exists r : B A with fr = gr = 1B. Note that FGFA𝜀FAFαFA is reflexive, with common right inverse FAFηAFGFA.
  • (b)
    By a split coequaliser diagram, we mean a diagram
    fghts A       B      C
    satisfying hf = hg, hs = 1C, gt = 1B and ft = sh. If these hold, then h is a coequaliser of (f,g) since if BkD satisfies kf = kg then k = kgt = kft = ksh, so k factors through h, and the factorisation is unique since h is (split) epic. Note that any functor preserves split coequalisers.
  • (c)
    Given G : D C, we say a pair AgfB in D is G-split if there’s a split coequaliser diagram
    GGhtsfg GA       GB      C
    in C. The pair (Fα,𝜀FA) in Lemma 5.9 is G-split, since
    GGαηηFGA𝜀FFαAA GF GF A     GF A     A
    is a split coequaliser diagram in C.

Theorem 5.11 (Precise Monadicity Theorem). A functor G : D C is monadic if and only if G has a left adjoint and creates coequaliser of G-split pairs in D.

Theorem 5.12 (Crude Monadicity Theorem). Assuming that:

Then G is monadic.

Proof.

Remark 5.13.

Example 5.14.

Definition 5.15 (Monadic tower). Let CGFD be an adjunction where D has reflexive coequalisers. The monadic tower of (FG) is the diagram

           ...

           (𝒞𝕋)𝕊


  𝒟        𝒞𝕋


GKLF          𝒞
where 𝕋 is the monad induced by (FG), and K and L are as in Theorem 5.7 and Lemma 5.9, and 𝕊 is the monad induced by (LK) and so on.

We say (FG) has monadic length n if we reach an equivalence after n steps.

6 Filtered Colimits

Definition 6.1 (Filtered). We say a category C is filtered if every finite diagram D : J C has a cone under it.

Lemma 6.2. C is filtered if and only if:

Proof.

For preorders, we say directed instead of filtered.

Definition 6.3 (Has filtered colimits). We say C has filtered colimits if every D : J C, where J is small and filtered, has a colimit.

Note that direct limits as in Example 4.3(g) are directed colimits.

Lemma 6.4. Assuming that:

Then C has all small colimits.

Proof. By Proposition 4.4(i), enough to show C has all small coproducts.

Given a set-indexeud family (Aj|j J) of objects, the finite coproducts jFAj, for F J finite, form the vertices of a diagram of shape PfJ = {F J|Ffinite} whose edges are coprojections. PfJ is directed, and a colimit for this diagram has the universal property of a coproduct jJAj.

Suppose given a D : I × J C, where C has limits of shape I and colimits of shape J.

    L(j)         L(j′)         colimJ L

                                          limI M


    D (i,j)       D(i,j′)                   M (i)


LDDDMD(β(i(α(α(α(i),β,,)′,D)j)j′β))(i′,j)       D(i′,j′)                   M (i′)
We can form L(j) = lim I(D(,j) : I C), by Example 4.7(e) these are the vertices of a diagram L : J C, and we can form colim JL.

Similarly, the colimits M(i) = colim JD(i,) form a diagram of shape I, and we can form lim IM. We get an induced morphism colim JL lim IM; if this is an isomorphism for all D : I × J C, we say colimits of shape J commute with limits of shape I in C.

Equivalently, colim J : [J,C] C preserves limits of shape I, or lim I : [I,C] C preserves colimits of shape J.

In Remark 5.13(d) we saw that reflexive coequalisers commute with finite products in Set.

Theorem 6.5. Assuming that:

Then colimits of shape J commute with all finite limits in Set if and only if J is filtered.

Proof.

Corollary 6.6. Assuming that:

Then

Proof.

Similar results hold for categories such as Cat.

Example 6.7. Consider the diagram

  ⋅⋅⋅     ℕ       ℕ      ℕ


sss111 ⋅⋅⋅     1       1      1
of shape op ×2 in Set. The inverse limit of the top row is , but that of the bottom row is 1. So lim op[op,Set] Set doesn’t preserve epimorphisms; equivalently colim : [,Setop] Setop doesn’t preserve monomorphisms. Thus by Remark 4.8, directed colimits don’t commute with pullbacks in Setop.

Given a functor F : C Set, the category of elements of F is (1F): its objects are pairs (A,x) with x FA and morphisms (A,x) (B,y) are morphisms f : A B such that (Ff)(x) = y.

Proposition 6.8. Assuming that:

Then the following are equivalent:

Proof.

Given a category C with filtered colimits, we say F : C D is finitary if it preserves filtered colimits. If C = Set, then a finitary F is determined by its restriction to Setf, since any set is the directed union of its finite subsets.

In fact the restriction functor [Set,D] [Setf,D] has a left adjoint (the left Kan extension functor) and the finitary functors are those in the image of this left adjoint (up to isomorphism).

For a category C as in Example 5.14(a) or Corollary 6.6, the corresponding monad 𝕋 on Set is finitary. From now on, Setf will denote the skeleton of the category of finite sets whose objects are the sets [n] = {1,2,,n}.

Definition 6.9 (Lawvere theory). By a Lawvere theory, we mean a small category T together with a functor Setf T which is bijective on objects and preserves finite coproducts. A model of a Lawvere theory T in any category C with finite products is a functor M : Top C preserving finite products.

For example, if 𝕋 is a monad on Set, the full subcategory of Set𝕋 whose objects are the sets [n] is a Lawvere theory.

Lemma 6.10. Assuming that:

Then the category of T-models in Set is (equivalent to) a finitary algebra category in the sense of Example 5.14(a).

Proof. Given a model M : Top Set, we have M[n]M[1]n for all n. Also, any morphism M[1]n M[1]p induced by a morphism [p] [n] in T is determined by its composites with the projections M[1]p M[1], so specifying M on morphisms is determined by its effect on morphisms with domain [1].

So, given a set A, specifying a model M with M[1] = A is equivalent to specifying operations αA : An A for each α : [1] [n] in T, subject to (vi)A(a1,,an) = ai whenever vi : [1] [n] is the i-th coprojection, and

  Ap      An


((γAαβA1)A,...,(βn)AA)
commutes whenever
  [1]      [n]


ασ(β1,...,βn)   [p]
commutes.

Note that the characterisation of T-models in any category with finite products. Note also that the equations of Lemma 6.10 allow us to reduce any compound operation α(β1(x),β2(x),,βn(x)) to a single operation γ.

Theorem 6.11. Assuming that:

Then the following are equivalent:
  • (i)
    C is equivalent to a finitary algebraic category in the sense of Example 5.14(a). (Recall that these categories are those whose objects are sets with finitary operations satisfying some equations, and morphisms are homomorphisms between them.)
  • (ii)
    C is equivalent to the category of Set-models of a Lawvere theory.
  • (iii)
    C Set𝕋 for a finitary monad 𝕋 on Set.

Proof.

For a general monad 𝕋 on Set, this construction produces a finitary monad 𝕋 which is the coreflection of 𝕋 in the category of finitary monads.

For example:

7 Regular Categories

Definition 7.1 (Image, cover). We say a category C has images if, for every AfB in C, there exists a least m : BB in Sub (B) through which f factors. We call m the image of f, and we say f is a cover if its image is 1B.

We write AfB to indicate that f is a cover.

Lemma 7.2. Any strong epimorphism is a cover. The converse holds if C has equalisers and pullbacks.

Proof. If f is strong epic, applying the definition to commutative squares of the form

 A       B ′


gfm1BB       B
shows that f is a cover.

For the converse, a cover AfB is epic since it can’t factor through the equaliser of any BhgC with gh. To verify the other condition, suppose given

 A       C

gfmhB       D
then the pullback of m along h is monic by Lemma 4.15, and f factors through it, so it’s an isomorphism. So we get B C by composing with the top edge of the pullback square.

Here, if C has images, image facorisation defines a functor [2,C] [3,C]: given

  A      B


fghk C      D
if we form the image factorisations
A      I       B


C      J       D
we get a unique I J making both squares commute.

Definition 7.3 (Regular category). We say C is regular if it has finite limits and images, and image factorisations are stable under pullback, i.e. if the left hand square above is a pullback then so are both right hand squares. (This is equivalent to saying that covers are stable under pullback).

Example 7.4.

Proposition 7.5. Assuming that:

Then covers coincide with regular epimorphisms.

Proof.

By a relation A B in a category C with finite products, we mean an isomorphism class of subobjects RA × B.

If C has images, we define the composite of ARBSC by forming the pullback

  P      S       C

  R      B


qpdcba A
forming the image of (ap,dq) : P A × C.

This is well-defined up to isomorphism and has the A(1A,1A)A × A as 2-sided identities.

Lemma 7.6. Composition of relations in C is associative if and only if C is regular.

Proof.

We write Rel(C) for the category whose objects are those of C and whose morphisms are relations. Note that Rel(Set) is just Rel as defined in Example 1.3(e).

We have a faithful functor C Rel(C) which is the identity on objects and sends AfB to A(1,f)A × B (for faithfulness, see Exercise 4.22(i)). We write f for (1A,f).

Note that there’s an isomporphism Rel(C) Rel(Cop) which is the identity on objects and sends R(a,b)A × B to R(b,a)B × A; we denote this by R, and write f for (f).

Also, Rel(C) is enriched over Poset (provided Rel(C) is locally small, i.e. C is well-powered), i.e. each Rel(C)(A,B) has a partial order which is preserved by composition.

We say ARB is left adjoint to BSA if 1A S R and R S 1B.

Proposition 7.7. ARB is a left adjoint in Rel(C) if and only if it is of the form f.

Proof.

P. Freyd developed a theory of allegories which have the structure of categories of relations and axiomatised those allegories A for which the subcategory Ala is regular.

In a regular category C, we say a relation R : AA is reflexive if 𝟙A R, symmetric if R = R, and transitive if R R R. R is an equivalence relation if it has all three properties. For any AfB in C, the kernel-pair R(a,b)A × A of f is an equivalence relation. We say an equivalence relation R is effective if it occurs as a kernel-pair, and C is effective regular if all equivalence relations are effective.

tfAbGp is regular but not effective regular: {(m,n) × |m n(mod2)} is a non-effective equivalence relation on .

Note that an equivalence relation is idempotent in Rel(C), and if A is an allegory and E is a class of symmetric idempotents in A then A[Eˇ] (as defined in Exercise 1.18) is an allegory; and if A is Rel(C) for a regular category C, then:

Proposition 7.8. Assuming that:

Then Ceff = (Rel(C)[Eˇ])la is effective regular, and the embedding Rel(C) Rel(C)[Eˇ] restricts to a full and faithful regular functor C Ceff which is universal among regular functors C D where D is effective regular.

Note that if C is effective regular, its equivalence relations are split idempotents in Rel(C): if ARA is the kernel-pair of AfB then it splits as ff (as we saw for C = Set in Exercise 1.19).

Definition 7.9 (Topos). A topos is a regular category E for which the embedding E Rel(E) sending f to f has a right adjoint. We write the effect of the right adjoint on objects by APA, and the unit A PA as {}A, and the counit PAA as APA × A.

In Set, PA is the power-set of A, the unit is the mapping a{a} of Example 1.7(c), and A = {(A,a)|a A} PA × A.

Note that (isomorphism classes of) subobjects of A are in bijection with morphisms 1 PA. C. J. Mikkelses showed that any topos has finite colimits; we’ll give Bob Paré’s proof, which is much simpler.

Proposition 7.10. Assuming that:

Then there exists a monadic functor Eop E. In particular, Eop has finite colimits and if E has limits of shape J then it also has colimits of shape Jop.

Proof. We make the assignment APA into a functor P : E E and a functor P : Eop E: given f : A B, Pf : PA PB corresponds to the image of APA × A1 × fPA × B, and Pf corresponds to the pullback of

                 ∃
                  B

1×f P B × A      P B × B
Given CgPA corresponding to RC × A, (Pf)g corresponds to the image of RC × A1 × fC × B and similarly given SD × B, composing with Pf corresponds to pulling back along D × A1 × fD × B.

Given a pullback square

  D      A

hkfg B      C
in E,
  P D      P A

PPPPk∗f∗Phg B      P C
commutes, since both ways correspond to the image of the left vertical composite in
E            ∃A


P A × D      P A × A

P A × B      P A × C
where both squares are pullbacks.

Now, as in Example 5.14(d), we have that if EeAgfB is a coreflexive in E, then

PPPPP∗∗∗gePfge B       PA       PE
is a split coequaliser coequaliser in E. Also, P is self-adjoint on the right, and it reflects isomorphisms by Exercise 7.17(v). The second assertion follows from Proposition 5.8(i).

Definition 7.11 (Support, totally supported, capital).

  • (a)
    By the support of an object A in a regular category, we mean the image of A 1. We say A is well-supported if A 1 is a cover.
  • (b)
    We say a regular category C is totally supported if every object is well-supported. We say C is almost totally supported if every object is either well-supported or a strict initial object, where we cann an object 0 strict if every A 0 is an isomorphism. (Given finite limits, a strict object is initial since for any A there exists 0π10 × Aπ2A, and the equaliser of any pair 0A is a).
  • (c)
    We say a regular category C is capital if its terminal object 1 is a detector, i.e. C(1,) reflects isomorphisms.

Example. Gp and AbGp are totally-supported since their terminal objects are initial. Set is almost totally-supported and capital. Note that capital implies almost totally-supported since if A isn’t well-supported there are no morphisms 1 A.

A representable functor C(A,) always preserves limits, so it’s a regular functor if and only if A is cover-projective (c.f. Definition 2.10).

Lemma 7.12. Assuming that:

Then 1 is cover-projective.

Proof. Since covers are stable under pullback, we need to show that every A1 is split epic. If A1, nothing to prove. If not, the projections A × AA aren’t equal (since their coequaliser is A1, by Proposition 7.5). So there exists 1 A × A not factoring through their equaliser, so there exists 1 A × A A.

If C is regular, the full subcategory Cws of well-supported objects is closed under finite products since

A × B     A


B         1
is a pullback, and under pullbacks of covers since if AB then A and B have the same support.

We write Ctv for the category obtained from Cws by adjoining a strict initial object 0: this is regular and almost totally-supported and the functor C Ctv sending all non-well-supported objects to 0 is regular (c.f. Exercise 5.19).

Lemma 7.13. Assuming that:

Then there exists an isomorphism-reflecting regular functor I : C C, where C is also small and almost totally-supported, such that for every well-supported A ob C there exists a morphism 1 IA in C not factoring through I(m) for any proper subobject m : AA in C.

Proof. Recall from Exercise 7.17: C regular implies CA regular for any A, and for any f : A B in C pullback along f defines a regular functor f : CB CA, which has a left adjoint Σf : CA CB sending g : C A to fg. And f reflects isomorphisms if and only if f is a cover.

We’ll define C as (C^)tv where C^ is easier to describe.

To satisfy the desired conclusion for a single well-supported object A, enough to take (!)A : CC1 CA, since (!A)A = (A × Aπ2A) acquires a point Δ : (A1A) (A × A A) not factoring through (A× A A) for any proper AA.

More generally, for any finite list A1,,An of well-supported objects, we can take C i=1nAi.

We define a base to be a finite list A = (A1,,An) of distinct well-supported objecs of C. We preorder the set B of bases by A B if B contains all the members of A. We write A for the product i=1nAi and if A B we write πB,A for the product projection B A. This makes A A into a functor Bop C.

Hence the assignment A C A, πB,AπB,A is ‘almost’ a functor B Cat.

We now define C^: its objects are pairs (B,f) where B is a base and f : A B is an object of C B. Morphisms (B,f) (B,f) are represented by pairs (C,g) where C is a base containing B and B and g : πf πf in C C, subject to the relation which identifies (C,g) with (C,g) if C C and the pullback of g to C C is isomorphic to g.

Clearly, each C B sits inside C^ as a non-full subcategory; so in particular CC [] is a subcategory of C^, C^ is regular, and the inclusions C B C^ are isomorphism-reflecting regular functors.

Given a finite diagram in C^, we can choose B such that all edges of the diagram appear as morphisms in C B, and take the limit there, and this is a limit in C^. Similarly for images.

Also, if a morphism f becomes an isomorphism in C^, its inverse must live C B for some B, hence f is an isomorphism C B.

We define C = (C^)tv: the induced functor C C^ C is still isomorphism reflecting since C is almost totally-supported.

Lemma 7.14. Assuming that:

Then there exists an isomorphism reflecting regular functor C C^ where C is capital. Hence in particular, there is an isomorphism-reflecting regular functor C Set.

Proof. Consider the sequence

C = C0 C1 C2 ,

where each Cn+1 is obtained from Cn by the construction of Lemma 7.13.

We define C^ to be the pseudo-colimit of this sequence: objects are pairs (n,A) where A ob Cn, and morphisms (n,A) (m,B) are represented by pairs (p,f) where p max {m,n} and F : IA IB in Cp, modulo the identification of (p,f) with (p,f) if p p and f = If.

The proof that C is regular, and that the embeddings Cn C^ are isomorphism-reflecting regular functors, is as in Lemma 7.13.

Given any non-invertible monomorphism AA in C^, it lives in Cn for some n, so there exists 1 A in Cn+1 not factoring through AA.

But if AfB isn’t monic in C^, the legs RbaA of its kernel-pair aren’t equal, so there exists 1rR not factoring through their equation, so 1brarA are distinct but have the same composite with f.

So C^(1,) reflects monomorphisms and hence reflects isomorphisms.

Theorem 7.15. Assuming that:

Then there exists a set I and an isomorphism-reflecting regular functor C SetI.

Proof. Let I be a representative set of subobjects of 1 in C, and for each U I consider the composite

C (!U) C U (CU)tv (CU)tv^ Set,

where the third factor is the functor of Lemma 7.14 and the fourth is represented by 1.

Given any non-invertible morphism AfB in C, if U is the support of B then (!)Hf remains non-invertible in CU and its codomain is well-supported there, so it remains non-invertible in (CU)tv and hence in Set.

So these functors collectively reflect isomorphisms.

Remark 7.16.

˙

Index

Eilenberg-Moore algebra

Lawvere theory

adjoint

adjunction

allegory

adjoint on the right

almost totally-supported

balanced

capital

category

categorical coproduct

categorical property

categorical coproduct

cartesian closed

coequaliser

colimit

complete

completeness

cone

contravariant

coproduct

cospan

counit

cover

cover-projective

creates

detector

detecting

diagram

directed

direct limit

direct sequence

edge

effective

effective regular

Eilenberg-Moore

(4)

(5)

(6)

epimorphism

epic

equaliser

equivalence

equivalent

equivalence relation

essentially injective

essentially surjective

faithful

filtered

full

functor

functorial

G-split

has filtered colimits

image

initial

inverse limit

inverse sequence

initial object

Kleisli category

Kleisli

left adjoint

commute

limit

locally small

model

monad

monadic

monic

monomorphism

(1)

(2)

(3)

natural

natural isomorphism

naturality square

natural transformation

naturality

naturally isomorphic

injective

reflexive

preserve

preserves

product

projective

pullback

pushout

pointwise

right adjoint

reflect

reflects

reflective

reflection

reflexive

reflective subcategory

regular

regular

regular epi

relation

representation

represented

representable

separator

separating

strict initial object

skeletal

skeleton

small

span

split coequaliser

split

solution-set condition

solution-set

subobject

support

symmetric

terminal

terminal object

topos

transitive

totally-supported

unit

vertex

well-copowered

weakly

well-powered

well-supported

Yoneda embedding